Topotactic Oxidation Pathway of ScTiO3 and High-Temperature

Joey A. Lussier , Kevin M. Szkop , Arzoo Z. Sharma , Christopher R. Wiebe , and Mario ... Joey A. Lussier , Graham Devitt , Kevin M. Szkop , Mario Bie...
0 downloads 0 Views 4MB Size
Article pubs.acs.org/IC

Topotactic Oxidation Pathway of ScTiO3 and High-Temperature Structure Evolution of ScTiO3.5 and Sc4Ti3O12-Type Phases∥ Shahid P. Shafi,† Bradley C. Hernden,† Lachlan M. D. Cranswick,‡ Thomas C. Hansen,§ and Mario Bieringer*,† †

Department of Chemistry, University of Manitoba, Winnipeg, Manitoba, Canada, R3T 2N2 Canadian Neutron Beam Centre, National Research Council Canada, Chalk River Laboratories, Chalk River, Ontario, Canada, K0J 1J0 § Institut Laue Langevine, 6 rue Jules Horowitz, BP 156, 38042 Grenoble Cedex 9, France ‡

S Supporting Information *

ABSTRACT: The novel oxide defect fluorite phase ScTiO3.5 is formed during the topotactic oxidation of ScTiO3 bixbyite. We report the oxidation pathway of ScTiO3 and structure evolution of ScTiO3.5, Sc4Ti3O12, and related scandium-deficient phases as well as high-temperature phase transitions between room temperature and 1300 °C using in-situ X-ray diffraction. We provide the first detailed powder neutron diffraction study for ScTiO3. ScTiO3 crystallizes in the cubic bixbyite structure in space group Ia3̅ (206) with a = 9.7099(4) Å. The topotactic oxidation product ScTiO3.5 crystallizes in an oxide defect fluorite structure in space group Fm3̅m (225) with a = 4.89199(5) Å. Thermogravimetric and differential thermal analysis experiments combined with in-situ X-ray powder diffraction studies illustrate a complex sequence of a topotactic oxidation pathway, phase segregation, and ion ordering at high temperatures. The optimized bulk synthesis for phase pure ScTiO3.5 is presented. In contrast to the vanadiumbased defect fluorite phases AVO3.5+x (A = Sc, In) the novel titanium analogue ScTiO3.5 is stable over a wide temperature range. Above 950 °C ScTiO3.5 undergoes decomposition with the final products being Sc4Ti3O12 and TiO2. Simultaneous Rietveld refinements against powder X-ray and neutron diffraction data showed that Sc4Ti3O12 also exists in the defect fluorite structure in space group Fm3̅m (225) with a = 4.90077(4) Å. Sc4Ti3O12 undergoes partial reduction in CO/Ar atmosphere to form Sc4Ti3O11.69(2).

1. INTRODUCTION Investigations of formation pathways of inorganic solids play a vital role in deducing structure−property relationships. In contrast to molecular sciences, reaction mechanisms are not well understood for extended solids. The recent advent of laboratory in-situ diffraction techniques allows investigation of solid-state reactions in real time. AVO3 phases with trivalent A cations (A = Ln, Sc, In, Y) are known to form either perovskites or bixbyites.1−7 The ideal cubic perovskite structure described in space group Pm3̅m (Figure 1a) has a 12-fold-coordinated large A cation and 6-foldcoordinated smaller B cation. The B−O6 octahedra form a corner-sharing infinite 3-dimensional network. Perovskite phases can undergo a number of distortions and cooperative tilts in order to accommodate differently sized A cations effectively. In contrast, the cubic bixbyite structure (Figure 1) is comprised of edge- and corner-sharing (A/B)−O6 octahedra with statistical A and B cation disorder. Notably, the bixbyite structure accommodates the cations on 2 distinct sites (8b and 24d). The 8b site forms a regular octahedron, whereas the 24d site is at the center of a distorted octahedron. Perovskite structures can be predicted with the Goldschmidt tolerance factor, Gt = (A−O)/√2(B−O), where A−O and B−O © 2012 American Chemical Society

correspond to the respective metal−oxygen bond distances. Perovskite phases are expected for Gt values between 0.8 and 1.1.8 For smaller A cations (e.g., Sc3+ and In3+) with tolerance factors less than 0.8 the AVO3 compounds may crystallize in the cation-disordered bixbyite structure (Figure 1b). We are interested in understanding solid-state reaction pathways with the potential benefit of controlling structures during synthesis. Understanding reaction sequences and solidstate reactivity will eventually provide insights into reaction mechanisms. Topotactic reactions involve modifications of structures while retaining substantial atomic connectivity of the precursor structures. Consequently, topotactic reactions do not require full reconstruction of the product crystal lattice. A recent review by Ranmohotti et al. discussed topochemical manipulations of perovskites in detail.9 Investigation of bixbyite phases and the structurally related fluorite structures with varying concentrations of oxide defects is motivated by their potential applications as solid-state oxide ion conductors.22 Controlling the oxide defect concentrations and defect structures as well as evaluation of defect structure Received: May 16, 2011 Published: January 24, 2012 1269

dx.doi.org/10.1021/ic201034x | Inorg. Chem. 2012, 51, 1269−1277

Inorganic Chemistry

Article

Figure 1. (a) Ideal cubic ABO3 perovskite structure with 12-fold-coordinated A cation (blue sphere) and 6-fold-coordinated B cation (yellow spheres). (b) ABO3 cubic bixbyite structure with A and B cation disorder arranged in (A/B)−O6 regular (green) and distorted (blue) octahedral environment. Oxygen atoms are represented as red spheres in both structures.

stabilities is a crucial ingredient for the design of reliable highperformance oxide ion conductors for use in solid oxide fuel cells. In 2004, Alonso et al. reported that the bixbyite ScVO3 undergoes topotactic oxidation to form the defect fluorite ScVO3.5 under relatively mild conditions.4 Shafi et al.10 showed that further oxidation of ScVO3.5+x results in the metastable intermediate ScVO4−x defect zircon structure prior to the topotactic oxidation to the fully oxidized ScVO4 zircon phase. In addition, our work on the oxidation of InVO3 bixbyite showed a similar metastable intermediate defect fluorite InVO3.5+x (0.00 ≤ x ≤ 0.22).11 The structural relationship between bixbyite and defect fluorite phases has been discussed in detail.11,12 Thus far, bixbyite oxidation pathway analysis has only been reported for vanadium-based compounds. We are investigating the oxidation of ScTiO3 bixbyite13 in search of the previously unknown ScTiO3.5 defect fluorite structure. The titanates provide a unique opportunity to stabilize this phase at high temperature because of the absence of higher oxidation states than 4+ for titanium. The ScTiO3 oxidation pathway will enable us to generalize the oxidative pathway of bixbyite phases. In addition, a large concentration of oxide defects can be maintained at high temperatures in air for systematic investigation of oxide defects for oxide ion conductor applications. Our investigation of the ScTiO3 oxidation pathway via in-situ X-ray diffraction covers the entire phase diagram. To the best of our knowledge, very few phases are known in the Sc−Ti−O phase diagram, namely, ScTiO3,13 Sc2TiO5,14 (ScxTi1−x)2O3,15 Sc4Ti3O12,16,17 Sc9Ti10O31.2,18 and Sc2Ti2O7.19−21 We are reporting for the first time the novel oxygen-deficient phases of ScTiO3.5 and Sc4Ti3O12−x. As the maximum oxidation state for Ti is 4+, the oxygen stoichiometry in ScTiO3.5 is restricted unlike in other AVO3.5+x (A = Sc, In) defect fluorite structures.

2Sc2O3 + 3TiO2 + Ti → 4ScTiO3

(1)

ScTiO3 + 0.25O2 → ScTiO3.5

(2)

Polycrystalline Sc4Ti3O12 was prepared by heating stoichiometric amounts of Sc2O3 (Alfa Aesar, 99.99%) and TiO2 (Alfa Aesar, 99.995%) at 1500 °C in air for 12 h according to eq 3. Sc4Ti3O12 was reduced in CO/Ar (1:3 ratio) flow in a tube furnace at 1500 °C, resulting in formation of the oxide defect phase Sc4Ti3O12−x according to eq 4.

2Sc2O3 + 3TiO2 → Sc4Ti3O12

(3)

Sc4Ti3O12 + xCO → Sc4Ti3O12 − x + xCO2

(4)

2.2. Room-Temperature Powder X-ray Diffraction. All products were identified using a PANalytical X’Pert Pro powder X-ray diffractometer with Cu Kα1,2 (λ = 1.540598 Å, 1.544426 Å) radiation equipped with a diffracted beam Ni filter and an X’Celerator detector operated in Bragg−Brentano geometry. The room-temperature diffractograms were collected from 10° to 120° in 2θ with a step width of 0.0083°. Phase identification (using PDF2003), preliminary indexing, and space group determination were carried out with X’Pert Highscore Plus (version 2.1). The powder X-ray diffraction data sets were analyzed in detail by the Rietveld method using FullProf 2008.23 2.3. High-Temperature Powder X-ray Diffraction. Hightemperature powder X-ray diffraction experiments were carried out on a PANalytical X’Pert Pro diffractometer equipped with a diffracted beam Ni filter, an X’celerator detector, and an Anton Paar HTK2000 high-temperature camera. Polycrystalline ScTiO3 was heated on a resistive platinum strip heater from 25 to 1300 °C in air. Using Cu Kα1,2 (λ = 1.540598 Å, 1.544426 Å) radiation diffraction patterns were collected in 20 °C increments covering the angular range 15° ≤ 2θ ≤ 120° with a 0.0167° step size. Using the same conditions thermal expansion experiments were conducted for ScTiO3.5 and Sc4Ti3O12 from 25 to 1300 °C at 50 °C increments in air. 2.4. Multiphase High-Temperature in-Situ Rietveld Refinements. The Rietveld refinements during heating of ScTiO3 in air were conducted with typically 28−37 parameters for 1, 2, and 3 phases depending on the temperature range as indicated below.

2. EXPERIMENTAL SECTION

(i) 25 °C ≤ T ≤ 220 °C: 1 phase, ScTiO3 cubic bixbyite phase (ii) 240 °C ≤ T ≤ 300 °C: 2 phases, ScTiO3 cubic bixbyite and ScTiO3.5 defect fluorite phases (iii) 320 °C ≤ T ≤ 840 °C: 1 phase, ScTiO3.5 defect fluorite phase (iv) 860 °C ≤ T ≤ 920 °C: 2 phases, Sc-rich and Ti-rich cubic fluorite defect phases

2.1. Synthesis. Polycrystalline ScTiO3 was prepared by conventional solid-state synthesis. Stoichiometric amounts of Sc2O3 (Alfa Aesar, 99.99%), TiO2 (Alfa Aesar, 99.995%), and Ti metal (Alfa Aesar, 99.99%), according to eq 1, were ground in an agate mortar with acetone and pelletized. The pellet was heated for 4−5 h in a diffusion pump vacuum (p < 10−4 mbar) at 1500 °C with one intermediate grinding using a Huettinger TIG 5/300 induction furnace equipped with a copper coil. Bulk ScTiO3.5 (∼500 mg) was synthesized according to eq 2 by heating ScTiO3 at 800 °C for 5 h in air. 1270

dx.doi.org/10.1021/ic201034x | Inorg. Chem. 2012, 51, 1269−1277

Inorganic Chemistry

Article

Figure 2. In-situ X-ray diffraction contour plot ((a) 21° ≤ 2θ ≤ 38° and (b) 61° ≤ 2θ ≤ 64°) of ScTiO3 oxidation in air from 25 to 1300 °C at 20 °C increments. Diffraction peak intensities are shown as constant increment contours from blue (lowest intensity) to red (highest intensity).

(v) 940 °C ≤ T ≤ 1060 °C: 3 phases, Sc-rich and Ti-rich cubic defect fluorite phases and TiO2 rutile (vi) 1080 °C ≤ T ≤ 1240 °C: 3 phases, cubic and rhombohedral Sc4Ti3O12 phases and TiO2 rutile (vii) 1260 °C ≤ T ≤ 1300 °C: 2 phases, cubic Sc4Ti3O12 defect fluorite phase and TiO2 rutile.

by powder X-ray and neutron diffraction. Our previous work on topotactic oxidations of bixbyite phases10,11 prompted our in-situ powder diffraction study of the oxidation pathway of ScTiO3. Figure 2 ((a), 21° ≤ 2θ ≤ 38° and (b), 61° ≤ 2θ ≤ 64°) is the contour plot of the temperature-dependent powder X-ray diffractograms during ScTiO3 oxidation in air. The contour plot shows ScTiO3 persisting up to approximately 300 °C; the onset of new peaks belonging to an intermediate phase at around 240 °C indicate oxidation of ScTiO3. This intermediate phase is stable up to 840 °C. The diffractograms of this novel phase resemble that of ScVO3.54 and InVO3.511 and can be indexed on a cubic unit cell with space group Fm3̅m. In analogy to ScVO3.5 and InVO3.5 the novel intermediate is an oxygen-deficient fluorite phase with composition ScTiO3.5. The low oxidation temperature (250− 300 °C) suggests that oxidation proceeds topotactically. This is also in agreement with the close structural relation between the bixbyite and the fluorite structure.4,11,12 Above 920 °C, ScTiO3.5 undergoes decomposition into TiO2 (rutile) and another scandium titanate phase consistent with space group Fm3̅m. The diffraction pattern of the defect fluorite phase observed at T ≥ 940 °C matches Sc4Ti3O12 (ICDD powder X-ray diffraction reference code 00-031-1227). At 1080 °C ≤ T ≤ 1240 °C the appearance of new peaks can be observed in Figure 2a and very prominently in Figure 2b. These additional peaks can be indexed on a rhombohedral unit cell in R3̅ space group and are consistent with the anion-ordered Sc4Ti3O12 structure.16 Since the anion-ordered (rhombohedral) and cation-disordered (cubic) Sc4Ti3O12 phases coexist between T = 1080 and 1240 °C, the disorder−order phase transition is first order. 3.1.1. Unit Cell Dimension Evolution During Oxidation and Annealing. Figure 3(top) shows the cubic unit cell volume evolution of ScyTizO3+x as a function of temperature during insitu oxidation of ScTiO3. The initial linear volume increase is due to thermal expansion of ScTiO3 (blue solid circles) followed by an up-turn, indicating oxygen uptake until 300 °C has been reached. From 250 °C the unit cell volume of the oxygen defect fluorite structure ScTiO3.5−x (red solid circles) increases until all Ti3+ has been oxidized to Ti4+. The concave rather than linear volume evolution between 400 and 700 °C is possibly due to oxide defect randomization in the ScTiO3.5 structure. The thermal expansion of fully characterized (see below) ScTiO3.5 and Sc4Ti3O12 were determined in two

For all data sets the unit cell parameters, scale factors, peak shape parameters, temperature factors, background parameters, and sample height were refined. No atomic positions were refined for the hightemperature structures. The χ2 values for all but two refinements were consistently between 2.0 and 2.9. More details are provided in the Supporting Information. 2.5. Room-Temperature Powder Neutron Diffraction. Roomtemperature powder neutron diffraction data for ScTiO3 were collected on the medium-resolution 800-wire diffractometer C2 operated by the National Research Council Canada in Chalk River. The diffraction patterns were measured with neutron wavelengths λ = 2.37248(15) Å (4° ≤ 2θ ≤ 84°) and 1.33076(10) Å (35° ≤ 2θ ≤ 115°) with 0.1° step sizes. Room-temperature powder neutron diffraction data for ScTiO3.5 and Sc4Ti3O12 were collected on the high flux diffractometer D2024 at the Institut Laue-Langevin (ILL) in Grenoble with λ = 1.8674(3) Å (germanium (115) reflection with a monochromator takeoff angle of 118°). Simultaneous powder neutron and powder X-ray refinements were carried out with FullProf 2008.23 2.6. Thermogravimetric Analysis/Differential Thermal Analysis (TGA/DTA). Simultaneous thermogravimetric and differential thermal analysis (TGA/DTA) experiments were carried out with a Linseis L81 thermobalance. ScTiO3 (approximately 43 mg) was heated from 25 to 1450 °C and then cooled to room temperature in static air at a rate of 20 °C/min. Experiments were conducted in alumina crucibles with an empty crucible as the reference. The buoyancy correction was carried out with an empty crucible in the sample position. Sc4Ti3O12−x (approximately 49 mg) was heated from 25 to 1400 °C at 20 °C/min in oxygen flow and then cooled to room temperature at 60 °C/min. Experiments were corrected for buoyancy and conducted in alumina crucibles with Al2O3 as the reference. All products were identified by powder X-ray diffraction.

3. RESULTS AND DISCUSSION 3.1. ScTiO3 Bixbyite Oxidation Pathway, ScTiO3.5/ Sc4Ti3O12 Formation, and High-Temperature Order− Disorder Transitions. ScTiO3 crystallizes in the reported13 cubic bixbyite structure in space group Ia3̅ (206) with scandium and titanium disorder on the 8b and 24d sites. The unit cell parameter a = 9.7099(4) Å is in agreement with the previously reported value of a = 9.709(3) Å by Reid et al.13 Below we report the detailed structure of ScTiO3 as determined 1271

dx.doi.org/10.1021/ic201034x | Inorg. Chem. 2012, 51, 1269−1277

Inorganic Chemistry

Article

phase (assigned as Sc3Ti4O12.5) in a continuous fashion until it reaches the composition Sc4Ti3O12. Loss of TiO2 increases the effective cationic radius on the 4a site due to the Sc3+ to Ti4+ ratio increase which is observable by the increase in slope of the unit cell volume (red solid circles to green solid rectangles). Thermal expansion of Sc4Ti3O12 (green solid line) is in very good agreement with this increased slope (green solid rectangles), and hence, the composition at high temperature during the in-situ study can be assigned as Sc4Ti3O12. The bottom panel in Figure 3 represents the temperature ranges for various observed phases during oxidation of ScTiO3. The Tirich regime is represented as Sc3Ti4O12.5 and the Sc-rich regime as Sc4Ti3O12 for the temperature range T = 860−1060 °C for convenience in Figure 3(bottom). A small (peak height 800 °C ScTiO3.5 decomposes into TiO2 rutile and cubic Sc4Ti3O12. We suggest that the driving force behind the decomposition of ScTiO3.5 is the thermodynamic stability of the rutile phase. 3.2.3. Sc4Ti3O12 Structure. It is noteworthy that there are only two entries in the ICSD and PDF2003 corresponding to Sc4Ti3O12, namely, an anion ordered rhombohedral structure and a cation-disordered cubic defect fluorite structure. Diffraction patterns obtained at T ≥ 900 °C during in-situ

Figure 7. Rietveld plots for ScTiO3 room-temperature refinement. Powder X-ray diffraction data Cu Kα1,2 radiation and powder neutron diffractograms λ = (a) 1.3295(2) and (b) 2.3726(5) Å. Red circles = experimental data, black line = best fit, blue line = difference, black tick marks = Bragg positions.

are provided in Table 1. The refined occupancies indicate a site preference of Ti3+ for the 8b site and a small Sc3+ preference for the 24d site. The refined composition Sc0.969(6)Ti1.025(6)O3 is in excellent agreement with the nominal composition. The bond valences (determined with VaList25) confirm these cation preferences. There is no indication of cation ordering in ScTiO3. 3.2.2. ScTiO3.5 Structure. Room-temperature X-ray and neutron powder diffraction data were collected on a bulk sample of ScTiO3.5, and the Rietveld analysis was carried out with FullProf 2008.23 ScTiO3.5 crystallizes in the defect fluorite structure in space group Fm3m ̅ (225) with a = 4.89199(5) Å. For the 2 histogram refinements a total of 19 parameters including peak shape parameters, scale factors, neutron wavelength, unit cell parameter, temperature factors, cation occupancies, and zero points were refined. The backgrounds were fitted using a cubic spline during the initial cycles and fixed for subsequent cycles. The Rietveld plots for ScTiO3.5 refined in space group Fm3̅m are shown in Figure 8, and structural details are provided in Table 2. The Sc3+/Ti4+ cations occupy the 4a (0, 0, 0) site in a disordered fashion, and the 1274

dx.doi.org/10.1021/ic201034x | Inorg. Chem. 2012, 51, 1269−1277

Inorganic Chemistry

Article

Table 1. Structural Parameters, Bond Angles, and Average Bond Distances for ScTiO3 Bixbyite Phase (space group Ia3̅ (No. 206)) As Obtained from Rietveld Refinements against Three Neutron and One X-ray Diffraction Pattern Measured at Room Temperature composition unit cell content T (K) space group unit cell Sc/Ti (8b) (1/4, 1/4, 1/4)

Sc/Ti (24d) (x, 0, 1/4)

O (48e) (x, y, z)

a (Å) V (Å3) Biso (Å2) Occ. (Sc) Occ. (Ti) x/a Biso (Å2) Occ. (Sc) Occ. (Ti) x/a y/b z/c Biso (Å2) Occ. (O)

d(Sc(1)/Ti(1))−O (Å) d(Sc(2)/Ti(2))−O (Å)

ScTiO3 Sc15.5(1)Ti16.4(1)O48 295 Ia3̅(206) 9.7099(4) 915.47(7) 0.76(9) 0.281(5) 0.719(5) 0.9665(1) 0.66(5) 0.554(3) 0.446(3) 0.3905(3) 0.1553(2) 0.3812(3) 0.43(4) 1.000 (fixed)

average d(Sc(2)/Ti(2)−O bond valences BV, site occupancy based on BV (site occupancy from refinement)

Sc(1) Ti(1) Sc(2) Ti(2)

no. of params Z R values

XRDa (Rp, Rwp, χ2) NPD-1a (Rp, Rwp, χ2) NPD-2a (Rp, Rwp, χ2) NPD-3a (Rp, Rwp, χ2)

6 × 2.081(2) 2 × 2.143(2) 2 × 2.108(2) 2 × 2.040(2) 2.097(2) 3.348, 43% (0.28) 2.742, 57% (0.72) 3.226, 61% (0.55) 2.642, 39% (0.45) 38 16 4.39/5.99/3.94 2.07/2.80/4.78 2.18/2.95/4.64 2.24/2.97/2.62

Table 1. continued

a X-ray: Kα1,2, λ = 1.540598 Å, 1.544426 Å, 10° ≤ 2θ ≤ 90°, Δ2θ = 0.0167°, 4781 data points, weight in refinement = 0.25. NPD-1: λ = 1.3295(2) Å, 4.4° ≤ 2θ ≤ 84°, Δ2θ = 0.1003°, 795 data points, weight in refinement = 0.25. NPD-2: λ = 1.3295(2) Å, 35.4° ≤ 2θ ≤ 115°, Δ2θ = 0.1003°, 795 data points, weight in refinement = 0.25. NPD-3: λ = 2.3726(5) Å, 4.4° ≤ 2θ ≤ 84°, Δ2θ = 0.1003°, 795 data points, weight in refinement = 0.25.

Table 2. Structural Parameters, Bond Angles, and Average Bond Distances for ScTiO3.5 Oxygen-Deficient Cubic Fluorite Phase (space group Fm3̅m (No. 225)) as Obtained from Rietveld Refinements against One Neutron and One X-ray Diffraction Pattern Measured at Room Temperature composition unit cell content T (K) space group unit cell Sc/Ti (4a) (0, 0, 0)

O (8c) (1/4, 1/4, 1/4) d(Sc(1)/Ti(1))−O (Å) no. of params Z R values

a (Å) V (Å3) Biso (Å2) Occ. (Sc) Occ. (Ti) Biso (Å2) Occ.

XRDa (Rp, Rwp, χ2) NPD-1a (Rp, Rwp, χ2)

ScTiO3.5 Sc1.94(2)Ti2.05(2)O7.00 295 Fm3̅m (225) 4.89199(5) 117.073(2) 2.43(1) 0.487(5) 0.513(5) 5.15 (3) 0.875 8 × 2.118292(4) 19 2 2.90/4.30/11.6 3.18/4.42/3.04

Figure 9. Fluorite structure with Sc3+/Ti4+ cations in blue and O2− anions in red. Randomly chosen oxygen defect is shown in the tetrahedral environment as a yellow sphere.

Table 3. Sc4Ti3O12 exists in the defect fluorite structure in space group Fm3̅m with unit cell parameter a = 4.90077(4) Å which is larger than that of ScTiO3.5 due to more Sc3+ ions in the former phase (Shannon radii Sc3+(VI) = 0.745 Å, Ti4+(VI) = 0.605 Å). The cations Sc3+/Ti4+ occupy the 4a (0, 0, 0) site in a disordered fashion, and the oxide anion is located on the 8c (1/4, 1/4, 1/4) site with an occupancy of 0.88. The bond valences (determined with VaList25) indicate the presence of Sc3+ (BV = 3.8) and Ti4+ (BV = 3.5) on one site. The particularly large oxygen temperature factor of 6.91(9) Å2 is due to the large defect concentration in the structure. The composition calculated from the Rietveld refinement is Sc2.16(1)Ti1.84(1)O7.03(4) with Z = 2. 3.3. Partial Topotactic Reduction of ScTiO3.5 and Sc4Ti3O12. Under strongly reducing conditions both defect fluorite phases ScTiO3.5 (H2 gas, 700 °C) and Sc4Ti3O12 (CO gas, 1500 °C) showed partial reduction of Ti4+ ions indicated by the

a X-ray: Kα1,2, λ = 1.540598 Å, 1.544426 Å, 10° ≤ 2θ ≤ 120°, Δ2θ = 0.0083°, 13 158 data points, weight in refinement = 0.3. NPD-1: λ = 1.8671(2) Å, 0.1° ≤ 2θ ≤ 150.9°, Δ2θ = 0.1001°, 1508 data points, weight in refinement = 0.7.

oxidation of ScTiO3 matched the cubic defect fluorite phase Sc4Ti3O12. The structural characteristics of Sc4Ti3O12 were established via combined Rietveld refinements against powder X-ray and neutron diffraction data measured on bulk samples using FullProf 2008.23 A total of 20 parameters including peak shape parameters, scale factors, neutron wavelength, unit cell parameter, temperature factors, occupancies of cations, and zero points were refined. The Rietveld refinement plots are shown in Figure 10, and the structural details are provided in 1275

dx.doi.org/10.1021/ic201034x | Inorg. Chem. 2012, 51, 1269−1277

Inorganic Chemistry

Article

Figure 11. Powder X-ray diffraction patterns zoomed in to the (111) reflection of Sc4Ti3O12 (red), Sc4Ti3O12−x (green), and the mixture (blue). Sample colors of Sc4Ti3O12 and Sc4Ti3O12−x are compared on the righthand side.

Figure 10. Rietveld plots for cubic Sc4Ti3O12 at room temperature. Powder X-ray diffraction data, Cu Kα1,2 radiation: (a) powder neutron diffractogram (D20) with λ = 1.8668(1) Å. Red circles = experimental data, black line = best fit, blue line = difference, black tick marks = Bragg positions. Pt peaks originating from the sample container in the neutron data have been excluded from the refinement.

Table 3. Structural Parameters, Bond Angles, and Average Bond Distances for Sc4Ti3O12 Oxygen-Deficient Cubic Fluorite Phase (space group Fm3̅m (No. 225)) as Obtained from Rietveld Refinements against One Neutron and One X-ray Diffraction Pattern Measured at Room Temperature composition unit cell content T (K) space group unit cell Sc/Ti (4a) (0, 0, 0)

O (8c) (1/4, 1/4, 1/4) d(Sc(1)/Ti(1))−O (Å) no. of params Z R values

a (Å) V (Å3) Biso (Å2) Occ. (Sc) Occ. (Ti) Biso (Å2) Occ.

XRDa (Rp, Rwp, χ2) NPD-1a (Rp, Rwp, χ2)

Sc4Ti3O12 Sc2.16(1)Ti1.84(1)O7.03(4) 295 Fm3̅m (225) 4.90077(4) 117.705(1) 3.28(2) 0.540(3) 0.460(3) 6.91(9) 0.88(2) 8 × 2.12210(2) 20 2 3.10/4.69/12.0 2.30/3.57/4.71

Figure 12. TGA/DTA oxidation of Sc4Ti3O12−x in air from 25 to 1400 °C at a heating rate of 20 °C/min. Red and blue lines indicate TGA and DTA curves, respectively, the solid arrow indicates the mass gain observed from room temperature to 1100 °C, and the dashed line is only a guide to the eye.

X-ray: Kα1,2, λ = 1.540598 Å, 1.544426 Å, 10° ≤ 2θ ≤ 120°, Δ2θ = 0.0083°, 13 158 data points, weight in refinement = 0.3. NPD-1: λ = 1.8668(1) Å, 0.1° ≤ 2θ ≤ 150.9°, Δ2θ = 0.1001°, 1508 data points, weight in refinement = 0.7. a

mass gain. The mass gain of 0.97% during oxidation indicates the composition of this unreported phase to be Sc4Ti3O11.69(2). The fully oxidized Sc4Ti3O12 phase has a larger unit cell due to its larger oxygen content in comparison to the reduced Sc4Ti3O12−x phase. This trend is consistent with the redox behavior of the defect fluorite phases AVO3.5+x (A = Sc, In) where unit cell expansion was observed with increasing oxygen stoichiometry rather than cell contraction due to substitution of the larger V4+ with the smaller V5+ cation.11

sample color change as well as the peak shift in the diffraction pattern. X-ray data for the (111) reflection of Sc4Ti3O12 and Sc4Ti3O12−x and their respective sample colors are shown in Figure 11. These two materials are visibly distinguishable as Sc4Ti3O12 is pale yellow while Sc4Ti3O12−x is black. Both phases produce the same diffraction patterns with an obvious peak position shift. Sc4Ti3O12 peaks (red peaks) are shifted to lower angles compared to those of Sc4Ti3O12−x (green). Also, the XRD pattern of the mixture of these two phases (blue pattern shown as an insert in Figure 11) showed double peaks indicating that the phases have different unit cell dimensions. Figure 12 shows the TGA/DTA plot during oxidation of Sc4Ti3O12−x in oxygen with a single exotherm and a single-step

4. SUMMARY AND CONCLUSIONS We are reporting for the first time the synthesis, stability, and structure of the novel oxygen defect fluorite ScTiO3.5. Oxidation of ScTiO3 has been followed via in-situ X-ray diffraction and TGA/DTA experiments. The topotactic oxidation of the bixbyite phase ScTiO3 resulted in formation of the related oxide defect structure ScTiO3.5; this is an 1276

dx.doi.org/10.1021/ic201034x | Inorg. Chem. 2012, 51, 1269−1277

Inorganic Chemistry



extension of the previously published oxidation pathways of the vanadium-bearing bixbyite phases ScVO3 and InVO3 and therefore clearly illustrates a structure−reactivity relation. With this the present study emphasizes a generalized topotactic oxidation pathway for bixbyite phases. In addition to the initial topotactic oxidation step the present in-situ study has revealed complex phase equilibria between cubic ScTiO3.5, Sc3Ti4O12.5, and Sc4Ti3O12 phases and an oxide ordered rhombohedral defect fluorite phase. ScTiO3.5 crystallizes in the defect fluorite structure (space group Fm3̅m) with Sc3+/Ti4+ disorder on the 4a site and O2− anions occupying the 8c site with 1/8 disordered defects. Neutron diffraction data suggest possible oxide defect clustering in ScTiO3.5; the same has been reported for the vanadium analogues.10 Only prolonged annealing of ScTiO3.5 at 800 °C just below the initial formation temperature of Sc4Ti3O12 results in a phase-pure homogeneous ScTiO3.5 sample. Only at temperatures above 840 °C the cations become sufficiently mobile for phase separation into Sc-enriched as well as Sc-depleted cubic defect fluorite structures. Below that temperature the ion mobility and therefore the chemical reactivity appears to be limited to the oxide sublattice: the topotactic regime. The average coordination number of both cations in ScTiO3.5 is seven. It appears that the observed phase transitions and phase separations are driven by enabling octahedral Ti4+ coordination. This is the case for the rutile structure, which is found as a byproduct during Sc4Ti3O12 formation and during oxide ordering where the resulting rhombohedral structure also has an octahedral site, which is assumed to be exclusively occupied by Ti4+. Controlling the reactivity of sublattices and the corresponding ion mobilities in those oxide defect structures is particularly important for ion conductors and consequently for solid-state electrolytes in fuel cell applications. In-situ diffraction has matured into a powerful tool for tackling reaction pathway analysis for solid-state reactions and thus provides important information for the controlled synthesis of extended solids.



Article

REFERENCES

(1) Mizokawa, T.; Khomskii, D. I.; Sawatzky, G. A. Phys. Rev. B 1999, 60, 7309. (2) Ren, Y.; Palstra, T. T. M.; Khomskii, D. I.; Pellegrin, E.; Nugroho, A. A.; Menovsky, A. A.; Sawatzky, G. A. Nature 1998, 396, 441. (3) Nguyen, H. C.; Goodenough, J. B. Phys. Rev. B 1995, 52, 324. (4) Alonso, J. A.; Casais, M. T.; Martinez-Lope, M. J. Dalton Trans. 2004, 9, 1294. (5) Miyasaka, S.; Okimoto, Y.; Iwama, M.; Tokura, Y. Phys. Rev. B 2003, 68, 100406. (6) Lundgren, R. J.; Cranswick, L. M. D.; Bieringer, M. J. Solid State Chem. 2006, 179, 3599. (7) Lundgren, R. J.; Cranswick, L. M. D.; Bieringer, M. Chem. Mater. 2007, 19, 3945. (8) Rao, C. N. R.; Raveau, B. Transition Metal Oxides; John Wiley & Sons: New York, 1988. (9) Ranmohotti, K. G. S.; Josepha, E.; Choi, J.; Zhang, J.; Wiley, J. B. Adv. Mater. 2011, 23, 7309. (10) Shafi, S. P.; Kotyk, M. W.; Cranswick, L. M. D.; Michaelis, V. K.; Kroeker, S.; Bieringer, M. Inorg. Chem. 2009, 48, 10553. (11) Shafi, S. P.; Lundgren, R. J.; Cranswick, L. M. D.; Bieringer, M. J. Solid State Chem. 2007, 180, 3333. (12) Ambrosini, A.; Duarte, A.; Poeppelmeier, K. R.; Lane, M.; Kannewurf, C. R.; Mason, T. O. J. Solid State Chem. 2000, 153, 41. (13) Reid, A. F.; Sienko, M. J. Inorg. Chem. 1967, 6, 521. (14) Kolitsch, U.; Tillmanns, E. Acta Crystallogr. 2003, E59, i36. (15) Rice, C. E.; Robinson, W. R. J. Solid State Chem. 1977, 21, 155. (16) Rossell, H. J. J. Solid State Chem. 1976, 19, 103. (17) Lyashenko, L. P.; Shcherbakova, L. G.; Belov, D. A.; Knotko, A. V. Inorg. Mater. 2009, 45, 543. (18) Pokrovskii, B.; Isaeva, E. V. Russ. J. Inorg. Chem. 1977, 22, 1103. (19) Knop, O.; Brisse, F.; Castelliz, L. Can. J. Chem. 1969, 47, 971. (20) Brixner, L. H. Inorg. Chem. 1964, 3, 1065. (21) Komissarova, L. N.; Shatskii, V. M. Zh. Prikl. Khim. 1966, 39, 2211. (22) Goodenough, J. B. Nature 2000, 404, 821. (23) Rodríguez-Carvajal, J. FullProf. 2k, Vers. 4.40; 2008. (24) Hansen, T. C.; Henry, P. F.; Fischer, H. E.; Torregrossa, J.; Convert, P. Meas. Sci. Technol. 2008, 19, 034001. (25) Wills, A. S. VaList for Windows, Vers. 1.5.8; 1998.

ASSOCIATED CONTENT

S Supporting Information *

Details regarding the temperature-dependent Rietveld refinements against the powder X-ray diffraction data collected during in-situ ScTiO3 oxidation. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*Phone: (204) 474-6258. Fax: (204) 474-7608. E-mail: Mario_ [email protected]. Notes ∥

Deceased.



ACKNOWLEDGMENTS M.B. acknowledges NSERC and CFI for operating and infrastructure support. S.P.S. is thankful for graduate student support (UMGF) from the University of Manitoba. Dr. Holger Kleinke and Jackie Xu are acknowledged for synthesis of ScTiO3. We thank the Institut Laue-Langevin (ILL) for providing neutron beam time and technical support. We acknowledge the National Research Council−Canadian Neutron Beam Centre (NRC−CNBC) for access to the neutron diffractometer C2 and technical support. 1277

dx.doi.org/10.1021/ic201034x | Inorg. Chem. 2012, 51, 1269−1277